Many-body effects, orbital mixing and cyclotron resonance in bilayer graphene
Abstract
In a magnetic field bilayer graphene supports, at the lowest Landau level, eight characteristic zero-energy levels with an extra twofold degeneracy in Landau orbitals . They, under general one-body and many-body interactions, evolve into pseudo-zero-mode (PZM) levels. A close look is made into the detailed structure and characteristics of such PZM levels and cyclotron resonance they host, with full account taken of spin splitting, interlayer bias, weak electron-hole asymmetry and Coulomb interactions. It is pointed out that the PZM levels generally undergo orbital level mixing (in one valley or both valleys) as they are gradually filled with electrons and that an observation of interband cyclotron resonance over a finite range of filling provides a direct and sensitive probe for exploring many-body effects and orbital mixing in bilayer graphene.
I Introduction
Graphene supports as charge carriers massless Dirac fermions which display unique and fascinating electronic properties. Recently, increasing attention is directed to bilayers NMMKF ; MF and fewlayers of graphene, where the physics and applications of graphene become far richer, with, e.g., a tunable band gap MF ; OBSHR in bilayer graphene.
In a magnetic field, graphene leads to a relativistic” infinite tower of electron and hole Landau levels, with the lowest Landau level (LLL) consisting of a number of zero-energy levels. The presence of such zero-mode levels has a topological origin in the index of (the leading part of) the Dirac operators, or in the chiral anomaly NS . Monolayer graphene carries zero-mode levels. Bilayer graphene supports an octet of such levels, with an extra twofold degeneracy in Landau orbitals and . In real samples, topological zero-energy levels evolve, by an interplay of general spin and valley splittings and Coulomb interactions, into a variety of pseudo-zero-mode (PZM) levels, or broken-symmetry quantum Hall states, as discussed theoretically BCNM ; KSpzm ; BCLM ; CLBM ; CLPBM ; NL ; MAF . Meanwhile it was noted that orbital degeneracy is also lifted by Coulomb interactions alone KS_Ls : Quantum fluctuations of the valence band split, like the Lamb shift in the hydrogen atom, the and levels appreciably. The orbital degeneracy and its lifting by the orbital Lamb shift are new features specific to the LLL in fewlayer graphene. Experimentally, partial or full resolution of the eightfold degeneracy of the LLL in bilayer graphene has been observed in transport FMY ; ZCZJ ; WAFM ; MEM ; VJB and capacitance MFW ; KFLH ; HLZW measurements, but the detailed structure of the LLL remains yet to be clarified.
Graphene supports cyclotron resonance (CR) in a variety of channels, both intraband and interband AbF . Interband resonance is specific to Dirac electrons, with a number of competing resonance channels simultaneously activated over a certain range of the total Landau-level filling factor . CR provides a useful means to look into the many-body problem in graphene. In experiment CR has been explored for monolayer JHTWS ; DCNN ; HCJL and bilayer HJTS ; OFBB graphene. Many-body effects on CR IWF ; BM ; KS_CR ; RFG ; KS_CR_eh ; ST ; KS_CRnewGR ; SL were first detected rather indirectly by a comparison of some leading intraband and interband resonances JHTWS . Recently, Russell et al. RZTW have reported a direct signal of many-body effects on CR in high-mobility hBN-encapsulated monolayer graphene; see also ref. prepH . They observed significant variations of resonance energies over a finite range of filling factor under a fixed magnetic field ; the point lies in extracting genuine Coulombic effects as a function of while controlling the running of the velocity factor with . The data show a profile nearly symmetric in and suggest a small band gap HuntYYY due to weak coupling to the hBN substrate.
No such -dependent study of interband CR is yet available for bilayer graphene. The purpose of this paper is to examine, in anticipation of a future experiment, the detailed characteristics of the LLL in bilayer graphene, with full account taken of spin splitting, interlayer bias, weak electron-hole asymmetry and Coulomb interactions, and to show how they are detected by a -dependent and fixed- survey of interband channels of CR. Particular attention is paid to electron-hole (-) conjugation and valley-interchange operations which relate the electron and hole bands within a valley or between the two valleys. We clarify how they govern the Landau-level and CR spectra.
In Sec. II we review some basic features of the effective theory of bilayer graphene in a magnetic field. In Sec. III we consider Coulombic corrections to Landau levels and CR. Such corrections inevitably contain ultraviolet divergences and, in Sec. IV, we carry out renormalization to extract observable many-body contributions. In Secs. V and VI, we examine how the PZM level spectra evolve with filling, note the general presence of orbital mixing and discuss how the associated many-body effects are revealed by interband CR. Section VII is devoted to a summary and discussion.
II bilayer graphene
The electrons in bilayer graphene are described by four-component spinor fields on the four inequivalent sites and in the Bernal-stacked bottom and top layers of honeycomb lattices of carbon atoms. Their low-energy features are governed by the two Fermi points and in the Brillouin zone. The intralayer coupling eV is related to the Fermi velocity m/s (nm) in monolayer graphene. Interlayer hopping via the dimer coupling eV makes the spectra quasi-parabolic MF in the low-energy branches .
The bilayer Hamiltonian with and leads to - symmetric energy spectra. Infrared spectroscopy, however, detected weak asymmetry ZLBF ; LHJ_asym due to the -sublattice energy difference meV and the nonleading interlayer coupling .
The effective Hamiltonian with such intra- and inter-layer couplings is written as MF ; NCGP
(5) |
with , and . Here denotes the electron field in valley , with and referring to the associated sublattices; stands for an interlayer bias, which opens a tunable valley gap OBSHR ; MF . We ignore the effect of trigonal warping eV which, in a strong magnetic field, causes only a negligibly small level shift. is diagonal in the (suppressed) electron spin.
As a typical set of band parameters we adopt those by recent theoretical calculations JM ; fn ,
(6) |
which are also used in an experimental analysis HLZW .
Let us note that is unitarily equivalent to with the signs of reversed,
(7) |
with ; signifies setting in . The electron and hole bands are therefore intimately related within a valley.
The Hamiltonian in another valley is given by with , and acts on a spinor of the form . Actually, is unitarily equivalent to with the sign of reversed,
(8) |
with . In what follows we adopt for and simply pass to valley by reversing the sign of bias in valley- expressions. We also suppose, without loss of generality, that in valley ; thus refers to valley .
The unitary equivalence
(9) |
implies that the - conjugation symmetry is kept exact only for although, if , it is apparently broken within each valley. This symmetry analysis also holds in the presence of Coulomb interactions.
Let us place bilayer graphene in a strong uniform magnetic field normal to the sample plane; we set, in , with , and rescale so that , where denotes the magnetic length.
The eigenmodes of are labeled by integers and plane waves with momentum , and have the structure
(10) |
where only the orbital modes are shown using the standard harmonic-oscillator basis , with for . The coefficients for each are given by the normalized eigenvectors of the reduced Hamiltonian
(11) |
where
(12) |
is the cyclotron energy for monolayer graphene; , , and . Numerically, for the set of parameters in Eq. (6) at T,
(13) |
and for meV. In what follows, we employ this set (13) of parameters for numerical estimates at T.
The unitary equivalence in Eq. (9) reveals how the spectra and the associated eigenmodes change via - conjugation. Within each valley they read
(14) |
while, between the two valleys, they are related as
(15) |
Of our particular concern are the and modes. For , has an obvious eigenvalue and eigenvector
(16) |
For , has three solutions , with belonging to the higher branches.
The mode (with ) and the mode have zero energy for and deviate from zero energy as develop. These pseudo-zero modes (PZM) form the LLL in bilayer graphene. The eigenenergy and eigenvector of the mode are written as
(17) |
In what follows, we denote valley and - symmetry breaking collectively as or . As seen from the secular equation, the correction in is odd in , and hence the normalization factor is even in ,
(18) |
Numerically, at T, while scarcely depends on [for small ].
Interlayer bias , if nonzero, breaks valley symmetry and shifts the PZM levels oppositely () in the two valleys. Accordingly, are nearly degenerate in each valley and are ordered so that for .
For , has rank 4 and we denote the four branches of Landau-level spectra as (with ) so that the index reflects the sign of . Let us denote in units of ,
(19) |
For , the (dimensionless) spectra of the lower branches take the form
(20) |
where , and ; is the sign function. The full spectra, to first order in breaking , are written as
(21) |
Actually, - breaking , listed in Eq. (13), has a sizable effect and shifts all the spectra, except for , upwards appreciably [e.g., roughly by 10 % for ], as seen from Fig. 1, which, for later convenience, is presented in Sec. V. Unlike , depend on bias only weakly . The spectra of the higher branches show similar but partially different dependence on breaking .
Each is associated with a pair of electron and hole modes (and ). In contrast, the pseudo-zero modes and stand alone (per spin and valley) and are, in this sense, - selfconjugate, with in Eqs. (14) and (15). Thus and ; that is, for and consists of odd powers of breaking .
The Landau-level structure is made explicit by passing to the basis (with ) via the expansion , where refers to the Landau level index, to the spin and to the valley. It is tacitly understood that the sum is taken over the higher branches as well. The one-body Hamiltonian is written as
(22) |
The charge density with is thereby written as KS_CR
(23) |
where ; stands for the center coordinate with uncertainty . The charge operators obey the algebra GMP .
The coefficient matrix in valley is constructed from the eigenvectors ,
(24) | |||||
where
(25) |
for , and ; for or ; . In view of Eqs. (14) and (15), have the following property under - conjugation,
(26) |
For , are even functions of breaking . They actually take simple form
(27) |
in each valley, with .
The Coulomb interaction is written as
(28) |
with the potential , and the substrate dielectric constant ; ; denotes normal ordering. For simplicity, we ignore a small effect of interlayer separation. Here and from now on, we suppress summations over levels , valleys and spins , with the convention that the sum is taken over repeated indices. The one-body Hamiltonian is thereby written as
(29) |
Here the Zeeman term is introduced via the spin matrix .
III Coulombic corrections
In this section we study Coulombic contributions to the Landau-level and associated CR spectra. In graphene, unlike conventional quantum Hall systems, the electrons and holes are always subject to quantum fluctuations of the infinitely-deep filled valence band (or the Dirac sea), which, being strong, lead to ultraviolet divergences. One first has to handle them properly.
The Coulomb direct interaction leads to a divergent self-energy , which, as usual, is removed when a neutralizing background is taken into account. The exchange interaction, on the other hand, gives rise to corrections to level spectra of the form
(30) |
Here stands for the filling fraction of the level. The exchange interaction preserves the spin and valley . Accordingly, from now on, we suppress them and mainly refer to valley .
The self-energies involve a sum over infinitely many filled levels in the valence band. Their structure is clarified if one notes the completeness relation
(31) |
Actually, this follows from the fact KS_LWGS that are () unitary matrices that obey the composition law with . The half-infinite sum in is then cast in the form
(32) |
where . Here again it is tacitly understood that the sum is taken over or as well.
The self-energies are thereby rewritten as
(33) |
Here the last term with the electron-hole” filling factor,
(34) | |||||
where for and otherwise, etc., refers to a finite number of filled electron or hole levels around the PZM sector .
Now involve a sum over infinitely many levels. for and yield ultraviolet divergences upon integration over with . In view of Eq. (26), are related in each valley or between the valleys,
(35) |
For , ; and are odd in breaking . The PZM sector thus has no divergence for , .
Let us eliminate from a constant, common to all levels, which is safely done by adjusting zero of energy. Via such regularization, self-energies are cast in the form
(36) |
Note that
(37) |
inherit the conjugation property of in Eq. (35).
A key property of the electron-hole filling factor defined in Eq. (34) is that it is odd under - conjugation: changes sign upon interchanging the electron and hole levels by replacing, in , . Noting this feature and rewriting Eq. (36) then allows us to relate, as done earlier KS_CRnewGR for monolayer graphene, in a valley or between the valleys. In terms of the full spectra to ,
(38) |
the result is
(39) |
Here specifies filling of the relevant valley; note that depend on filling. We assign to the empty PZM sector (of a given spin and valley) with levels below it () all filled, or with the uppermost filled level” ; accordingly, , e.g., refer to . refers to the PZM sector with one filled level and to the filled sector. For definiteness, in what follows, we focus on cases of integer filling, in which a distinct band gap is present, and take in Eq. (39) to be integers. Let us now imagine, e.g., valley with . Interchanging electrons and holes (according to ) yields a configuration with , and Eq. (39) implies that the spectra of such - conjugate configurations are intimately related. A typical - conjugate pair are the empty PZM sector () and the filled one ().
Some care is needed in handling the configuration, which is not uniquely fixed, e.g., with or or with any linear combination of . Note also that - conjugation reverses the level layout of the PZM sector, e.g., . As a result, if, e.g., in a valley, Eq. (39) reads
(40) |
which means that the filled level turns into the filled level in the conjugate configuration. In Sec. V, we study the PZM sector over a continuous range and encounter an interesting case of mixed levels at .
Let us next study CR, namely, optical interlevel transitions at zero momentum transfer, with the selection rule AbF for graphene, i.e., (i) intraband channels and (ii) interband channels for ; , , , etc. Interband CR is specific to Dirac electrons. Consider now CR from level to level for each (valley, spin)= channel and denote the associated excitation energy as ; CR preserves the valley and spin so that will be suppressed from now on. In the mean-field treatment KS_CR ; KH ; MOG , the corrections
(41) |
consist of the self-energy difference and the Coulombic attraction between the excited electron and created hole. The full CR spectra thus differ from the dressed level gaps by attraction energies ,
(42) |
It will be clear now that, via - conjugation, turns into . In particular, the interband CR channels are intimately related,
(43) | |||||
we suppose here. Such an - conjugate” character of was noticed earlier KS_CRnewGR for monolayer graphene, for which - conjugation is an exact symmetry. Here for bilayer graphene - conjugation, though not a symmetry, is an operation that tells us how the CR spectra deviate from the symmetric ones by breaking . Experimentally the conjugate channels of a given set are observable as competing signals over a finite range of filling factor and are indistinguishable unless polarized light is used.
IV renormalization
The self-energies are afflicted with ultraviolet divergences. In this section we consider how to extract physically observable information out of them. To this end, one first defines renormalized parameters by setting , , , etc., with counterterms , , etc. The one-body Hamiltonian is then divided into the portion involving only the renormalized quantities and the counterterms . One starts with , calculates the Coulombic corrections and encounters divergences. If one could remove them by adjusting , renormalization is done properly.
Fortunately, the renormalization procedure to for bilayer graphene in a magnetic field has been formulated earlier KS_CR_eh . The key point is that, since simply acts as a long-wavelength cutoff , the short-distance structure of the present theory is known from its free-space () version: (i) The divergence associated with velocity is the same as that for monolayer graphene,
(44) |
with a momentum cutoff . (ii) and remain finite,
(45) |
We thus take and finite; . (iii) and , though mixed under renormalization, are also governed by ,
(46) |
where .
Let us now rewrite the bare level spectra as . The renormalized spectra are the portion consisting of , , , etc. The associated counterterms , to be written as with finite coefficients , are expressed in terms of and are uniquely fixed once one specifies (or, the divergence ) by referring to a specific observable quantity; see Appendix A for details. The dressed level spectra, rewritten as
(47) |
thereby reveal observable self-energies , with the divergences in removed by . Also the full CR energies in Eq. (42) take a renormalized form,
(48) |
with corrections free from divergences.
Let us here define by referring to CR in the channel, as chosen experimentally HJTS . We fix so that takes a naive form ; i.e., . A neat way to handle this prescription is to replace in [of Eq. (36)] by
(49) |
and cast in the form
(50) | |||||
(51) |
Here , with
(52) |
is a kernel associated with . In this renormalized form, are sizable only for and vanish rapidly as ; one can thus calculate many-body corrections numerically without handling divergences. Note that we have carried out renormalization in such a way that and hence also share the same conjugation properties as in Eq. (35), i.e.,
(53) |
Once the eigenmodes are fixed numerically for a given set of parameters , one can now construct , , and and calculate the spectra and . Table I shows a list of of our interest, in units of
(54) |
For the PZM levels, and are practically linear in and are barely modified by - breaking ; indeed, . For other levels - breaking is evident, .
V Orbital mixing and Cyclotron resonance within the PZM sector
In bilayer graphene the LLL consists of sets of nearly degenerate levels. For the empty PZM sector [of a given (spin, valley)] at one can explicitly write down the renormalized spectra by substituting into Eq. (50),
(55) |
with . Here and from now on all quantities refer to renormalized ones. In this section we use to denote the spectra, .
In Eq. (55), the first terms are odd in breaking while the last terms are even in . For , only the latter remain. The orbital degeneracy is therefore lifted by quantum corrections even for zero breaking , giving rise to the orbital Lamb shift KS_Ls , with lower than by
(56) |
Numerically, at T, , and , or meV for a choice of or . The dressed spectra considerably deviate from the one-body spectra , as shown in Fig. 1 for and T at bias meV.

Let us write, for , the full shift as
(57) |
Numerically, at T,
(58) |
Obviously, for . The term in , , thus tends to reduce the orbital Lamb shift, e.g., for meV at T. At the same time, makes slightly less dependent on than . As a result, the nearly degenerate PZM levels , being quasi linear in , necessarily have a crossing or level inversion in either valley (per spin) as bias is varied.

These features are indeed seen from Fig. 2(a), which shows how such level spectra (per spin) change with bias for ; solid curves refer to valley and long-dashed curves to ; . Level inversion takes place across , or numerically, across or meV. Such a critical bias gets smaller as is made weaker and for higher , as seen from Figs. 2(b) and 2(c); lies in valley while lies in valley . Level inversion is also present for , with meV at T.
When the PZM sector is gradually filled with electrons, those level spectra come down (via exchange interaction) and, when the sector is filled up, they turn into the ones, depicted with thin dashed and dotted curves in Fig. 2. These spectra are - conjugate to the spectra,
(59) |
and are given by
(60) |
At , orbital shift is reversed in the layout and is considerably enhanced by , as is clear from Fig. 2. Incidentally, for , these conjugate spectra look symmetric about the axis. In this sense, - breaking is seen as an apparent asymmetry in Fig. 2. [Actually, asymmetry is particularly sizable for the level; with , one finds , and () for and .]
Let us now consider what will happen when we pass from to in valley with bias kept fixed in the range . It is the level that starts to be filled, getting lower in energy. The level also follows but, when is reached, it gets even lower than the level. This signals a level crossing or instability with filling, which actually is avoided via mixing of levels KS_Ls . For no such mixing takes place. Orbital mixing is therefore triggered by level inversion in the PZM spectra. (i) When , orbital mixing arises in both valleys for and in one valley () for . (ii) When , mixing takes place only in one valley () for , as seen from Fig. 2(b). (Here we suppose no level inversion in the spectra and take, e.g., meV.)
To see how level mixing proceeds let us rotate to in the orbital space,
(61) |
where and ; we refer to the levels associated with as . We also define the rotated charges by , where for , for and , etc.; are written with and with .
The PZM sector for is now governed by the one-body Hamiltonian [with the spectra for ] plus Coulomb interaction [with in Eq. (28)] acting within this sector. One can readily diagonalize it using the Hartree-Fock approximation: The rotated levels have the spectra KS_Ls ,
(62) |
where , and
(63) |
denote the filling fractions of the levels, with , etc. Note that enjoy the reciprocal relation
(64) |
Diagonalization of the PZM spectra is achieved for that obeys the relation
(65) | |||||
i.e., or , where .

Figure 3(a) depicts how the spectra change with filling factor for and T. For critical changes arise at three points with , and orbital mixing takes place in the range : (i) The level first gets filled. deviates from zero at , and changes as
(66) |
until is reached at . (ii) Filling of the level then starts, and changes according to
(67) |
attains or at and ceases to change. Via mixing the levels are interchanged,
(68) |
Actually, noting Eq. (65), one can simplify the level spectra when is moving, i.e., for ,
(69) |
The spectra then obey the relation
(70) |
which implies that, via - conjugation, the filled level turns into the filled within a valley. The small gap at develops into a sizable orbital gap at , as seen from Fig. 3(a).
On the other hand, for , no such mixing arises, but show similar behavior with filling. At , correctly obey the relation in Eq. (40) and have again a sizable gap.
In bilayer graphene the empty LLL (at ) consists of four sets of empty PZM sectors. Let us now consider how the LLL is filled over the range . As an illustration we focus on some typical cases in which the eightfold degeneracy of the LLL is fully lifted. Suppose first that bias is chosen so that the valley gap dominates over the spin gap . Filling of the LLL will then starts in valley with a spin component of the lowest energy, and each {valley, spin} set of PZM sectors will be filled in the following order
(71) |
Figure 1 illustrates such a sequence of level spectra for . There the total filling factor increases with each valley-filling factor , and a sizable band gap emerges at each integer filling: (i) The and gaps are orbital gaps (in red) of width . (ii) The gaps are enhanced spin gaps (in purple), , associated with full filling of one valley. (iii) Most prominent is the gap, which is a valley (+ orbital) gap (in green), . As seen from Figs. 2(a) and 2(b), the gap and gaps are essentially the same at zero bias , but the former increases rapidly with while the latter barely change.
On the other hand, when the spin gap dominates over the valley gap in high field , i.e., , filling of the LLL will proceed in the following order
(72) |
and the gap will open as an enhanced spin gap. Two such filling sequences (71) and (72) appear consistent with part of the (-dependent) layer-charge pattern observed in a recent capacitance measurement KFLH .
Let us next consider CR supported by the PZM sector (per spin and valley). Consider a time-dependent uniform electric potential , coupled to the PZM sector via the currents and , with the Hamiltonian
(73) |
where , and . Accordingly, CR takes place in channels , , , etc.
Notably, intra PZM resonance is possible BCNM . When orbital mixing is present (), CR arises in the channel, with excitation energy
(74) |
where , and
(75) |
On the other hand, when orbital mixing is absent , the channel is activated, with
(76) |
Interestingly, for , there is no correction to CR since holds; see Eq. (27).
Figure 3(b) shows how the excitation spectra evolve with filling at bias , meV (thus in valley ) and meV . While a sizable level gap opens around , it is almost cancelled by Coulombic attraction, leaving and of magnitude of the Lamb shift or even smaller.
Such an intra-PZM channel of CR is activated when an orbital band gap develops around . The coupling to the charge, however, is weak. It will therefore be a challenge to detect CR within the LLL in experiment. On the contrary, CR from or into the LLL serves as a practical probe into the LLL, as discussed in the next section.
VI interband cyclotron resonance

VI.1 PZM, PZM}
Figure 4(a) shows how the level spectra evolve as the PZM sector is gradually filled over (per spin and valley). The PZM spectra are shifted almost uniformly with bias while the spectra are barely affected. As a result, the associated channels of CR sensitively depend on .
For orbital mixing is present in both valleys and CR arises in four channels, and over the range ; see Appendix B for details of the CR spectra. The resonance spectra change in composition and strength in a characteristic way with filling , as depicted in Fig. 4(b) for at and 20 T. The strength of each response (function) is proportional to , i.e, the filling-factor difference between the initial and final levels mixing weight read from in Eq. (73). In Fig. 4(b) we also plot such relative weights, for and for . In the spectra red circles indicate the most prominent signal at each integer filling.
The and channels remain active even for and , respectively, i.e., when either the or level is partially filled. The associated CR spectra slightly rise there, because Coulombic attraction diminishes as or ; the CR signals themselves also vanish in this limit.
When bias is increased, e.g., to , orbital mixing disappears in valley , with only the and channels of CR activated, while orbital mixing continues in valley , as depicted in Fig. 4(c).
From these spectra one can visualize how the channels (or more) of compete as the LLL is filled over the range under a given bias . Figure 4(e) illustrates such resonance spectra of meV at each integer filling, with the valley-dominant filling sequence of Eq. (71) assumed. Also Fig. 4(d) shows the spectra, with the spin-dominant filling sequence ( of Eq. (72) assumed. It is clear that, with increasing bias , the spectra get split in the valley. Remarkably, the resonance spectra look nearly - symmetric, i.e., symmetric in . An asymmetry is apparent at odd integers in Fig. 4(e), where orbital mixing only arises in valley .
It is illuminating here to look into some numbers. From Figs. 4(a)(c) one can read off the following set of level gaps and excitation energies,
(77) | |||||
(78) | |||||
(79) | |||||
(80) | |||||
(81) | |||||
(82) |
in units of meV. These level gaps and CR spectra at = (0,2) are related via - conjugation,
(83) |
In view of this, Eqs. (77) and (78) imply that the effect of - breaking is on the order of 1% in these level and CR spectra. This is further confirmed by the approximate equality between the meV expressions. It is somewhat surprising that the level gaps and CR spectra are only slightly affected by - breaking while the level spectra themselves are considerably modified (by a few % generally, and much more for ), as noted in a paragraph below Eq. (60). Accordingly, in experiment, the CR signals of will look nearly symmetric in at even-integer fillings. A shift or splitting of signals with around odd-integer filling and will reveal the presence or absence of orbital mixing.
It is seen from Eq. (77) that Coulombic attraction accounts for roughly 30% of the relevant level gap here for . It diminishes rapidly for higher ; it amounts to about 12% for and 8% for .

VI.2 Interband resonance

Let us next consider other interband CR, . Figure 5(a) shows how the competing spectra evolve with filling of the relevant valley at bias meV. Thin dotted lines guide the spectra of the case; they are symmetric about at zero bias . As - breaking is turned on, and deviate upward and downward, respectively, by roughly 2 % from the symmetric spectra, making the spectra split more on the side and less on the other.
Figure 5(b) illustrates how the channels of spectra evolve over the range at bias meV, with the valley-dominant filling sequence of Eq. (71) assumed. Splittings among the spectra, unlike , develop with noticeably, and - breaking makes the full spectra visibly asymmetric in , in contrast to the - symmetric spectra, depicted in Fig. 5(c).
Figure 6 shows the CR spectra of and per spin and valley. Included in the figure are the values of total filling factor to be realized when (except ) is common to all valleys and spins, i.e., . These higher-energy resonances are practically insensitive to the detailed structure of the LLL and to a bias of meV. Again the CR spectra are less affected by than level shifts , but are made visibly asymmetric in . The competing spectra overlap around or and split more and more as for and for .
Presumably, when changes over a wide range as in Fig. 6, screening of the Coulomb potential will become important, as noted KS_CRnewGR ; SL for monolayer graphene. Via screening will get weaker with increasing , making the spectra decrease faster for larger .
VII Summary and discussion
Characteristic to bilayer graphene in a magnetic field is an octet of PZM levels nearly degenerate in orbitals as well as in spins and valleys. In this paper we have studied some basic characteristics of such PZM levels and shown that they generally undergo mixing in orbitals as they are gradually filled with electrons. We have examined possible consequences of orbital mixing and how they are detected by an observation of some leading competing channels of interband CR over a finite range of filling factor .
It will be illuminating to summarize here why and how orbital mixing arises. Let us first suppose an - symmetric setting with and . Coulomb interactions then lead to the orbital Lamb shift. The resulting shift is odd under - conjugation and changes sign as one goes from the empty to filled PZM sector, ; this is because the zero-modes are - selfconjugate [so that for ]. It is this level inversion that drives orbital mixing with filling of the PZM sector, as discussed in Sec. V. When valley and - breaking is turned on, the level shift takes a modified form at , with , and begins to change with bias slightly and almost linearly. The bias thereby acquires a critical value , beyond which orbital mixing disappears. In this way, orbital mixing is driven by the orbital Lamb shift, and generally takes place in either valley or both (per spin) as bias is varied.
In our analysis, special attention has been paid to - conjugation and valley-interchange operations, which govern, as in Eqs. (39) and (43), the level and CR spectra in bilayer graphene. In experiment, Coulombic corrections will be seen as variations of the interband CR spectra with filling , - breaking as an asymmetry of the spectra about , and valley breaking as variations or splitting of the spectra with bias . Orbital mixing will be detected by an opening of some additional channels of CR. The channels of interband CR, in particular, will serve as a direct and most sensitive probe to explore the novel characteristics of the LLL in bilayer graphene.
Appendix A Counterterms
In Sec. IV the bare level spectra are written as . In this appendix we outline how to calculate the counterterms numerically. For given , one can write the associated counterterms as
(84) |
with the differential operator
(85) |
acting on and with defined in Eq. (46). This formula allows one to evaluate analytically. Alternatively, one can let act on the reduced matrix in Eq. (11), and write
(86) |
Actually, is given by with substitution first and subsequently and . In this way one can directly calculate from eigenmodes .
Appendix B Resonance Spectra
In this Appendix, we outline the derivation of the CR spectra examined in Sec. VI. Let us first write the renormalized level spectra as . As the PZM sector is gradually filled over the range , each level acquires an additional self-energy correction
(87) |
Direct calculation of and leads to in Eq. (62). Adding Coulombic attraction terms
(88) |
then yields the CR spectra .
When orbital mixing is present, i.e., for , interband CR, , hosts four active channels (per spin and valley) over the range . Associated with the level are the excitation spectra,
(89) | |||||
(90) |
and those associated with the level are
(91) | |||||
(92) |
In actual calculations one can simplify, for , , , etc., under symmetric integration . Evaluating Eqs. (89) (92) numerically leads to the excitation spectra depicted in Fig. 4.
References
- (1) E. McCann and V. I. Fal’ko, Phys. Rev. Lett. 96, 086805 (2006).
- (2) K. S. Novoselov, E. McCann, S. V. Morozov, V. I. Fal’ko, M. I. Katsnelson, U. Zeitler, D. Jiang, F. Schedin, and A. K. Geim, Nat. Phys. 2, 177 (2006).
- (3) T. Ohta, A. Bostwick, T. Seyller, K. Horn, and E. Rotenberg, Science 313, 951 (2006).
- (4) A. J. Niemi and G. W. Semenoff, Phys. Rev. Lett. 51, 2077 (1983).
- (5) Y. Barlas, R. Côté, K. Nomura, and A. H. MacDonald, Phys. Rev. Lett. 101, 097601 (2008).
- (6) K. Shizuya, Phys. Rev. B 79, 165402 (2009).
- (7) Y. Barlas, R. Côté, J. Lambert, and A. H. MacDonald, Phys. Rev. Lett. 104, 096802 (2010).
- (8) R. Côté, J. Lambert, Y. Barlas, and A. H. MacDonald, Phys. Rev. B 82, 035445 (2010).
- (9) R. Côté, W. Luo, B. Petrov, Y. Barlas, and A. H. MacDonald, Phys. Rev. B 82, 245307 (2010).
- (10) R. Nandkishore and L. Levitov, Phys. Rev. B 82, 115124 (2010).
- (11) M. Mucha-Kruczyski, I. L. Aleiner, and V. I. Fal’ko, Phys. Rev. B 84, 041404(R) (2011).
- (12) K. Shizuya, Phys. Rev. B 86, 045431 (2012).
- (13) B. E. Feldman, J. Martin and A. Yacoby, Nature Phys. 5, 889 (2009).
- (14) Y. Zhao, P. Cadden-Zimansky, Z. Jiang, and P. Kim, Phys. Rev. Lett. 104, 066801 (2010).
- (15) R. T. Weitz, M. T. Allen, B. E. Feldman, J. Martin and A. Yacoby, Science 330, 812 (2010).
- (16) A. S. Mayorov, D. C. Elias, M.Mucha-Kruczynski, R. V. Gorbachev, T. Tudorovskiy, A. Zhukov, S. V. Morozov, M. I. Katsnelson, V. I. Fal’ko, A. K. Geim, K. S. Novoselov, Science 333, 860 (2011).
- (17) J. Velasco Jr., et al., Nat. Nanotechnol. 7, 156 (2012).
- (18) J. Martin, B. E. Feldman, R. T. Weitz, M. T. Allen, and A. Yacoby, Phys. Rev. Lett. 105, 256806 (2010).
- (19) A. Kou, B. E. Feldman, A. J. Levin, B. I. Halperin, K. Watanabe, T. Taniguchi, and A. Yacoby, Science 345, 55 (2014).
- (20) B. M. Hunt, J. I. A. Li, A. A. Zibrov, L. Wang, T. Taniguchi, K. Watanabe, J. Hone, C. R. Dean, M. Zaletel, R. C. Ashoori, and A. F. Young, Nat. Commun. 8, 948 (2017).
- (21) D. S. L. Abergel and V. I. Fal’ko, Phys. Rev. B 75, 155430 (2007).
- (22) Z. Jiang, E. A. Henriksen, L. C. Tung, Y.-J. Wang, M. E. Schwartz, M. Y. Han, P. Kim, and H. L. Stormer, Phys. Rev. Lett. 98, 197403 (2007).
- (23) R. S. Deacon, K.-C. Chuang, R. J. Nicholas, K. S. Novoselov, and A. K. Geim, Phys. Rev. B 76, 081406(R) (2007).
- (24) E. A. Henriksen, P. Cadden-Zimansky, Z. Jiang, Z. Q. Li, L.-C. Tung, M. E. Schwartz, M. Takita, Y.-J. Wang, P. Kim, and H. L. Stormer, Phys. Rev. Lett. 104, 067404 (2010).
- (25) E. A. Henriksen, Z. Jiang, L.-C. Tung, M. E. Schwartz, M. Takita, Y.-J.Wang, P. Kim, and H. L. Stormer, Phys. Rev. Lett. 100, 087403 (2008).
- (26) M. Orlita, C. Faugeras, J. Borysiuk, J. M. Baranowski, W. Strupiński, M. Sprinkle, C. Berger, W. A. de Heer, D. M. Basko, G. Martinez, and M. Potemski, Phys. Rev. B 83, 125302 (2011).
- (27) A. Iyengar, J. Wang, H. A. Fertig, and L. Brey, Phys. Rev. B 75, 125430 (2007).
- (28) Yu. A. Bychkov, and G. Martinez, Phys. Rev. B 77, 125417 (2008).
- (29) K. Shizuya, Phys. Rev. B 81, 075407 (2010).
- (30) R. Roldán, J.-N. Fuchs, and M. O. Goerbig, Phys. Rev. B 82, 205418 (2010).
- (31) K. Shizuya, Phys. Rev. B 84, 075409 (2011).
- (32) J. Sári and C. Töke, Phys. Rev. B 87, 085432 (2013).
- (33) K. Shizuya, Phys. Rev. B 98, 115419 (2018).
- (34) A. A. Sokolik and Y. E. Lozovik, Phys. Rev. B 99, 085423 (2019).
- (35) B. J. Russell, B. Zhou, T. Taniguchi, K. Watanabe, and E. A. Henriksen, Phys. Rev. Lett. 120, 047401 (2018).
- (36) J. Pack, B. J. Russell, Y. Kapoor, J. Balgley, J. Ahlers, T. Taniguchi, K. Watanabe, and E. A. Henriksen, preprint arXiv:2001.08879 [cond-mat.mes-hall].
- (37) B. Hunt, J. D. Sanchez-Yamagishi, A. F. Young, M. Yankowitz, B. J. Leroy, K. Watanabe, T. Taniguchi, P. Moon, M. Koshino, P. Jarillo-Herrero, and R. C. Ashoori, Science 340, 1427 (2013).
- (38) L. M. Zhang, Z. Q. Li, D. N. Basov, and M. M. Fogler, Z. Hao, and M. C. Martin, Phys. Rev. B 78, 235408 (2008).
- (39) Z. Q. Li, E. A. Henriksen, Z. Jiang, Z. Hao, M. C. Martin, P. Kim, H. L. Stormer, andD. N. Basov, Phys. Rev. Lett. 102, 037403 (2009).
- (40) J. Nilsson, A. H. Castro Neto, F. Guinea, and N. M. R. Peres, Phys. Rev. B 78, 045405 (2008).
- (41) J. Jung and A. H. MacDonald, Phys. Rev. B 89, 035405 (2014).
- (42) In ref. KS_CR_eh we adopted a slightly different set of band parameters. All our numerical analysis in the present paper has also been made with it and essentially the same conclusion has been reached.
- (43) S. M. Girvin, A. H. MacDonald, and P. M. Platzman, Phys. Rev. B 33, 2481 (1986).
- (44) K. Shizuya, Int. J. Mod. Phys. B 33, 1950171 (2019).
- (45) C. Kallin and B. I. Halperin, Phys. Rev. B 30, 5655 (1984).
- (46) A. H. MacDonald, H. C. A. Oji, and S. M. Girvin, Phys. Rev. Lett. 55, 2208 (1985).